Unstructured Grid Techniques: D. J. Mavriplis
Unstructured Grid Techniques: D. J. Mavriplis
UNSTRUCTURED GRID
TECHNIQUES
Annu. Rev. Fluid Mech. 1997.29:473-514. Downloaded from arjournals.annualreviews.org
D. J. Mavriplis
Institute for Computer Applications in Science and Engineering NASA Langley
Research Center, Hampton, VA 23681-0001
by CAPES on 03/24/09. For personal use only.
ABSTRACT
An overview of the current state of the art in unstructured mesh techniques for
computational fluid dynamics is given. The topics of mesh generation and adapta-
tion, spatial discretization, and solution techniques for steady flows are covered.
Remaining difficulties in these areas are highlighted, and directions for future
work are outlined.
1. INTRODUCTION
The field of computational fluid dynamics has been evolving rapidly over the
past several decades. Aerospace applications, particularly in the field of aero-
dynamics, have provided the major impetus for the advances in computational
fluid dynamics technology. Great strides in available computational power,
coupled with rapid advances in algorithmic efficiency, have enabled computa-
tional fluid dynamics to substantially reduce the amount of wind-tunnel time
required for airframe and propulsion system design.
At the same time, computational methods are continuously being required
to deliver more accurate solutions for more complex realistic configurations at
lower computational cost. Traditionally, structured or block-structured meth-
ods that rely on regular arrays of quadrilateral or hexahedral cells in two or
three dimensions, respectively, have been employed to discretize the compu-
tational domain. Unstructured grid methods originally emerged as a viable
alternative to the structured or block-structured grid techniques for discretizing
complex geometries. These methods make use either of collections of simpli-
cial elements (triangles in two dimensions, tetrahedra in three dimensions) or
473
0066-4189/97/0115-0473$08.00
December 2, 1996 16:42 Annual Reviews Chapter14 AR23-14
474 MAVRIPLIS
of elements of mixed type with irregular connectivity. This not only provides
greater flexibility for discretizing complex domains but also enables straightfor-
ward implementation of adaptive meshing techniques where mesh points may
be added, deleted, or moved about, while mesh connectivity is updated locally,
in order to enhance solution accuracy.
The advent of powerful desktop computers has greatly expanded the appli-
cability of computational fluid dynamics outside the realm of pure research
Annu. Rev. Fluid Mech. 1997.29:473-514. Downloaded from arjournals.annualreviews.org
and aeronautics applications, into all areas of fluid mechanics, as well as into
multidisciplinary fields. The generality of unstructured grid methods and their
ability to enhance solution accuracy through adaptive procedures have proved
to be such a great advantage for these diverse applications that many if not most
commercial computational fluid dynamics codes currently rely on the use of
unstructured meshes.
by CAPES on 03/24/09. For personal use only.
Despite their recent success, unstructured mesh techniques still incur sub-
stantially more overhead than structured grid methods, a fact that has limited
their use particularly for large three-dimensional viscous flow cases. The devel-
opment of efficient and robust unstructured mesh algorithms for grid generation
and flow solution represents a considerable ongoing challenge. This paper pro-
vides an overview of the current state-of-the-art of such techniques; it points
out the advantages and deficiencies of various approaches as well as promising
avenues for future improvements.
In section 2 an overview of the most common grid generation and adaptivity
techniques is given. Section 3 describes the various approaches to spatial dis-
cretization for the Euler and Navier-Stokes equations, and section 4 describes
techniques for integrating the discretized equations to steady-state. Owing
to space limitations, the additional but equally important topics of unsteady
solution techniques and parallel implementations are not covered. Survey pa-
pers on these items can be found in the literature (see Special course 1992;
Venkatakrishnan 1995 for example).
2. MESH GENERATION
The initial phase of any numerical simulation begins with the generation of a
suitable mesh. The basic premise is that, owing to the random way elements
may be assembled to fill computational space, unstructured mesh generation
techniques are inherently more automatic and amenable to complex geometries
than traditional structured or block-structured grids. Therefore, the problem of
unstructured mesh generation is largely one of designing algorithms that are
automatic, robust, and yield suitable element shapes and distributions for the
flow solver.
December 2, 1996 16:42 Annual Reviews Chapter14 AR23-14
In the following section, we examine the most prevalent techniques for gen-
erating unstructured meshes: the advancing-front method, and Delaunay-based
approaches. While the advancing-front method is somewhat heuristic in nature,
Delaunay-based methods are firmly rooted in computational geometry princi-
ples. In practice, both methods have resulted in successful three-dimensional
grid generation tools for arbitrary geometries. All of these techniques have also
demonstrated robustness problems and grid quality issues that have led to the
Annu. Rev. Fluid Mech. 1997.29:473-514. Downloaded from arjournals.annualreviews.org
Figure 1 Generation of a new triangle using new point (a), or existing front point (b), in the
advancing-front method. Dashed line edges denote current front.
December 2, 1996 16:42 Annual Reviews Chapter14 AR23-14
476 MAVRIPLIS
the stack is empty, i.e. when all fronts have merged upon each other and the
domain is entirely covered. One of the critical features of such methods is the
placement of new points. Upon generating a new triangle, a new point is first
placed at a position that is determined to result in a triangle of optimal size
and shape. The parameters that define this optimum triangle as a function of
field position are obtained by a prescribed field function. (This field function
is often user defined and is evaluated by interpolating from values stored on a
Annu. Rev. Fluid Mech. 1997.29:473-514. Downloaded from arjournals.annualreviews.org
coarse background grid.) The triangle generated with this new point may result
in a cross-over with other front edges, and may therefore be rejected. This
is determined by computing possible intersections with “nearby” front edges.
Alternately, an existing point on the front may coincidentally be located very
close to the new point, and thus should be employed as the forming point for
the new triangle, to avoid the appearance of a triangle with a very small edge at
by CAPES on 03/24/09. For personal use only.
some later stage. Existing candidate points are thus also searched by locating
all “nearby” front points.
For three-dimensional grid generation, a surface grid is first constructed by
generating a two-dimensional triangular mesh on the surface boundaries of the
domain. This triangular mesh forms the initial front, which is then advanced into
the flow-field by placing new points ahead of the front and forming tetrahedral
elements. The required intersection checking now involves triangular front
faces rather than edges as in the two-dimensional case.
Advancing-front methods generally result in smooth high-quality triangula-
tions in most regions of the flow-field. However, difficulties may be encoun-
tered in regions where the fronts must be merged. The algorithm is somewhat
heuristic in nature, and the need to perform intersection checking operations
on the moving front requires the use of dynamic quad/octree data structures for
efficient implementations.
strategy where the mesh points have been predetermined. The mesh points are
put in a list and an initial triangulation is artificially constructed (with auxiliary
points) that completely covers the domain to be gridded. The mesh points in
the list are then inserted sequentially into the existing triangulation using the
Bowyer-Watson algorithm. The final mesh is obtained when all points from the
list have been inserted. The mesh point distribution can in principle be random,
but most often it is either pregenerated using Cartesian mesh stencils (Baker
1987) or multiple overlapping structured meshes (Weatherill 1985; Mavriplis
1990a, 1991b) or may be generated on the fly using automatic point placement
strategies.
A simple automatic point placement technique has been proposed in Holmes
& Snyder (1988). Starting with an initial coarse triangulation that covers the
entire domain, a priority queue is constructed based on some parameter of the
individual triangles (circumradius, for example). A field value is assumed to
exist that determines the local maximum permissible value for the circumradius
of the triangles (or any other parameter). The first triangle in the queue is
examined, and a point is added at its circumcenter if the triangle circumradius
is larger than the locally prescribed maximum. This new point is inserted
into the triangulation using Bowyer-Watson’s algorithm, and the newly formed
triangles are inserted into the queue if their circumcircles are too large; otherwise
they are labeled as acceptable and do not appear in the queue. The final grid is
obtained when the priority queue empties out.
Delaunay triangulation techniques based on point insertion extend naturally
to three dimensions by considering the circumsphere associated with a tetrahe-
dron. These methods are more efficient than advancing-front techniques owing
to the simplicity of the algorithm. However, they are only valid on convex
domains. Therefore, Delaunay triangulation methods typically employ an arti-
ficially constructed convex domain (such as a quadrilateral in two dimensions or
December 2, 1996 16:42 Annual Reviews Chapter14 AR23-14
478 MAVRIPLIS
Annu. Rev. Fluid Mech. 1997.29:473-514. Downloaded from arjournals.annualreviews.org
a cube in three dimensions) that covers the entire flow-field. The triangulation is
by CAPES on 03/24/09. For personal use only.
initially performed in this domain, and the triangles or tetrahedra that lie in the
flow-field are then extracted from the original convex domain to form the com-
putational mesh. The main disadvantages of Delaunay triangulation techniques
relate to their inability to guarantee boundary integrity. In two dimensions, the
set of edges that define the domain boundaries must form a subset of the edges
of the original convex domain triangulation. If this is not the case, edges of the
triangulation exist that penetrate the boundary, as shown in Figure 3.
In three dimensions, the equivalent problem is that of constructing a tetrahe-
dralization of the domain that contains, as a subset of its faces, a given surface
boundary triangulation. The use of local transformations as described in the
following section (edge and face-edge swapping, in two and three dimensions,
respectively), has been developed for locally modifying Delaunay meshes to
conform to specified boundary discretizations (George et al 1991, Weatherill
et al 1993).
In two dimensions, the existence of constrained Delaunay triangulations
(Chew 1989) guarantees the possibility of constructing triangulations that con-
tain a prescribed subset of edges and thus respect a given boundary discretiza-
tion. In three dimensions, the recovery of prescribed boundary triangulations
may not always be achievable through swapping techniques. A somewhat less
restrictive problem is that of constructing a tetrahedralization of the boundary
and interior points for which no edges or faces intersect the boundary surfaces.
One way of achieving this is by modifying the boundary point resolution (by
judiciously adding surface points) in regions where breakthroughs occur (Baker
1991).
2.3 Edge and Face Swapping Techniques
The Delaunay triangulation represents but one of many possible triangulations
of a given point set. An algorithm for transforming one triangulation to another
December 2, 1996 16:42 Annual Reviews Chapter14 AR23-14
d
a
a
c
c b
Annu. Rev. Fluid Mech. 1997.29:473-514. Downloaded from arjournals.annualreviews.org
2D e
3D
Figure 4 Two possible configurations for the diagonal of a convex pair of triangles in the edge-
by CAPES on 03/24/09. For personal use only.
swapping algorithm, and for two tetrahedra in the three-dimensional face-edge swapping algorithm.
triangulation of the same point set is given by Lawson (1972). Since all two-
dimensional planar graphs obey Euler’s formula (Mortenson 1985), all possible
triangulations of a given set of points contain the same number of edges and
triangles. Thus, any one triangulation may be obtained by simply rearranging
the edges of another triangulation of the same set of points. The edge-swapping
algorithm of Lawson (1972, 1977) consists of successively examining pairs of
neighboring triangles in the mesh. For each pair of triangles there exist two
possible configurations of the diagonal edge, as shown in Figure 4.
The algorithm chooses the diagonal configuration that optimizes some given
criterion. The algorithm terminates when no further optimizations are possible.
The edge-swapping algorithm can be used to transform an arbitrary triangula-
tion into a Delaunay triangulation. In this case, the edge-swapping criterion to
be used is the edge configuration that maximizes the minimum angle within the
two neighboring triangles. (Delaunay triangulations are also known as max-min
triangulations: They maximize the smallest angles in the triangulation.)
Edge-swapping can be used to construct triangulations other than the De-
launay triangulation, such as the minimum maximum angle (min-max) trian-
gulation, or the minimum total edge length triangulation (Preparata & Shamos
1985, Barth 1991a). However, the edge-swapping procedure is not guaranteed
to result in the global optimum triangulation for these cases: The algorithm
may terminate within local optima. A more sophisticated and complex algo-
rithm, (O(N 2 log N )), known as the edge-insertion algorithm (Edelsbrunner
et al 1992), has been developed for transforming a given triangulation into the
global optimum min-max triangulation.
In three dimensions, analogous local transformations can be achieved through
face- and edge-swapping. By considering pairs of neighboring tetrahedra, the
December 2, 1996 16:42 Annual Reviews Chapter14 AR23-14
480 MAVRIPLIS
common triangular face may be removed and the edge joining the two oppos-
ing end points of the tetrahedra can be drawn, as shown in Figure 4. Note that
the original configuration that contained two tetrahedra is transformed into a
configuration containing three tetrahedra. Joe (1989) has shown that, when a
new point is introduced into an existing Delaunay tetrahedralization and con-
nected in an arbitrary manner, the global Delaunay tetrahedralization of the new
point set can be recovered through a sequence of face-edge swappings. How-
Annu. Rev. Fluid Mech. 1997.29:473-514. Downloaded from arjournals.annualreviews.org
ever, the more general case of transforming an arbitrary tetrahedral mesh into
the Delaunay tetrahedral mesh through face-edge swapping often terminates
prematurely in a local optimum.
Face-edge swapping procedures are nevertheless useful in improving the
quality of three-dimensional meshes. In fact, Delaunay tetrahedralizations are
not necessarily the best constructions for three-dimensional meshes, since sliver
by CAPES on 03/24/09. For personal use only.
482 MAVRIPLIS
The drive towards full Navier-Stokes solvers has necessitated the development
of stretched grid generation techniques in order to resolve the thin boundary-
layers, wakes, and other viscous regions characteristic of high-Reynolds-
number viscous flows. Proper boundary-layer resolution usually requires mesh
spacing several orders of magnitude smaller in the direction normal to the
boundaries than in the streamwise direction, resulting in large cell aspect-ratios
in these regions.
In Babushka & Aziz (1976) it is shown how the accuracy of a two-dimensional
finite-element approximation on triangular elements degrades as the maximum
angle of the element increases. Therefore, stretched obtuse triangles that con-
tain one large angle and two small angles are to be avoided, while stretched
right-angle triangles, with one small and two nearly right angles, are preferred.
Delaunay triangulations, which maximize the minimum angles of any trian-
gulation, tend to produce equiangular triangulations and are thus ill suited for
the construction of highly stretched triangular elements. One of the earliest ap-
proaches for generating highly stretched triangulations for viscous flows makes
use of a Delaunay triangulation performed in a locally mapped space (Mavriplis
1990b, 1991b; Vallet et al 1991; Castro-Diaz et al 1995). By defining a mapped
space based on the desired amount and direction of stretching, an isotropic De-
launay triangulation can be generated in this mapped space that, when mapped
back to physical space, provides the desired stretched triangulation. Difficulties
with such methods involve defining the stretching transformations and deter-
mining suitable point distributions for avoiding obtuse triangular elements.
The fact that obtuse triangles should be avoided has led to the development
of methods based on minimum-maximum angle triangulations in both two and
three dimensions (Barth 1991a, Marcum 1995). These can be implemented ei-
ther as a transformation from a Delaunay or other triangulation to a more optimal
min-max triangulation or by generating the min-max connectivity on the fly,
December 2, 1996 16:42 Annual Reviews Chapter14 AR23-14
as points are inserted, using local reconnection techniques. Although the same
reconnection strategy can be employed in stretched and unstretched regions of
the flow-field, different point placement strategies must be employed in these
regions, in order to obtain the required anisotropic mesh-point resolution in
the stretched regions, and to avoid the formation of obtuse triangular elements,
which cannot be eliminated by reconnection strategies alone. Point-placement
strategies for the highly stretched regions most often mimic structured hyper-
Annu. Rev. Fluid Mech. 1997.29:473-514. Downloaded from arjournals.annualreviews.org
bolic mesh generation techniques, where the points are placed along normal
lines emanating from each boundary vertex at intervals determined by the spec-
ified degree of mesh stretching. One of the drawbacks of min-max triangulation
approaches is that the resulting connectivity may be very sensitive to the dis-
tribution of vertices, with very small differences in mesh-point distributions
leading to substantially different connectivities.
by CAPES on 03/24/09. For personal use only.
484 MAVRIPLIS
Annu. Rev. Fluid Mech. 1997.29:473-514. Downloaded from arjournals.annualreviews.org
by CAPES on 03/24/09. For personal use only.
486 MAVRIPLIS
are most important here, while de-refinement has only a minor effect and can
often be omitted for steady-state cases. For transient problems, mesh adapta-
tion must be performed every several time steps, and thus efficiency is much
more important than optimality. Mesh refinement and de-refinement are both
essential for transient cases, as well as mesh movement for cases with moving
boundaries. Furthermore, the accuracy of interpolation from the original mesh
to the refined mesh affects the solution accuracy (unlike the steady-state case),
and thus accurate transfer schemes are required.
Delaunay-based mesh generation techniques can easily be extended to in-
corporate adaptive refinement capabilities (Mavriplis 1990a, 1995b; Weatherill
et al 1993; Barth 1992). Once a solution has been obtained on an initial mesh,
new points can be added in regions where high errors are detected. These
new points can be triangulated into the mesh using the Bowyer-Watson point-
insertion algorithm. Alternatively, if a non-Delaunay mesh is employed, new
points may be inserted through element subdivision, and the connectivity of the
resulting mesh may be optimized through several face-edge swapping passes
based on any appropriate criterion.
Rule-based hierarchical element subdivision is a very effective adaptive tech-
nique, particularly for unsteady flows, where efficiency and accuracy of inter-
polation between successive grids are important considerations (Stoufflet et al
1987, Löhner & Baum 1992, Rausch et al 1992, Braaten & Connell 1996). The
essential approach consists in recursively subdividing mesh elements where
large solution errors are detected, but conforming to a set of well-defined sub-
division rules, which are necessary to prevent the formation of degenerate ele-
ment shapes and connectivities. Storing the hierarchy of the recursive process
enables de-refinement to be implemented by simply retracing the subdivision
history. Hierarchical subdivision is also applicable to nonsimplicial hybrid
meshes by constructing a library of subdivision rules for each element type.
December 2, 1996 16:42 Annual Reviews Chapter14 AR23-14
criterion represents the weakest point of most adaptive strategies. The main
problem is that an exact characterization of the error requires a knowledge of
the solution itself, which is obviously impractical. Most error estimates are
based on the assumption that the solution is smooth and asymptotically close
to the exact solution. This is often not the case for fluid dynamics problems,
which are governed by nonlinear hyperbolic partial-differential equations. So-
lutions may contain discontinuities, and downstream flow features often depend
on adequate resolution of upstream flow features. Most refinement criteria are
heuristically based on (undivided) gradients and/or second differences of the
various flow variables. Conservative criteria (i.e. over-refining) are often em-
ployed to compensate for the inability to accurately characterize the true solution
error. Because adaptive meshing results in different mesh topologies for each
simulation, even when the geometry of the problem is unchanged, parametric
studies (typically used in design processes) are complicated by the requirement
to distinguish between grid-induced and physical solution variations. Never-
theless, for problems with disparate length scales, adaptive meshing is often
indispensable for resolving small flow features, and their full potential awaits
the development of more well-founded adaptive criteria.
3. SPATIAL DISCRETIZATION
The previous section describes techniques for generating suitable meshes about
arbitrary geometric configurations. Once such a mesh has been generated, it
serves as the basis for the spatial discretization of the governing fluid dynamic
equations. Unstructured mesh discretization techniques can generally be clas-
sified as finite-volume or finite-element strategies, depending on whether a
discrete integral or variational viewpoint is adopted, although many common
discretizations may be simultaneously interpreted from both viewpoints.
December 2, 1996 16:42 Annual Reviews Chapter14 AR23-14
488 MAVRIPLIS
LR
Annu. Rev. Fluid Mech. 1997.29:473-514. Downloaded from arjournals.annualreviews.org
LR
by CAPES on 03/24/09. For personal use only.
Figure 7 Illustration of dual mesh for mixed triangular-quadrilateral unstructured mesh and as-
sociated control-volumes with edge-based fluxes for a vertex-based scheme.
a tetrahedral mesh five to six times more elements than vertices. On the dual
mesh, the degree of a vertex (number of incoming edges) is fixed and equal to
three for triangular elements, or four for tetrahedral elements, whereas on the
original mesh, the degree of each vertex is variable. Similar relationships hold
for other elements such as prisms and pyramids.
Thus, cell-centered and vertex-based schemes operating on the same grid
result in vastly different discretizations when nonquadrilateral or nonhexahedral
Annu. Rev. Fluid Mech. 1997.29:473-514. Downloaded from arjournals.annualreviews.org
evidence to suggest that they also provide more accurate solutions on equivalent
grids. The question of whether the additional overheads are offset by the
increase in accuracy is still an open one, especially since vertex-based schemes
operate on more complex stencils (more fluxes per unknown) than cell-centered
schemes.
3.1 Central Differencing and Riemann-Based
Upwind Methods
A simple finite-volume discretization can be obtained by defining a control
volume about each unknown (mesh-element for cell-centered schemes, dual-
mesh cell for vertex-based schemes) and integrating the fluxes over the boundary
of the control volume using a simple midpoint integration rule, where the value
of the flux at each control volume face, as shown by the arrows in Figure 7, is
computed by averaging the flow variables in the two control volumes on either
side of the face. For vertex-based schemes on simplicial meshes, this particular
scheme can also be derived using Galerkin finite elements, where the fluxes are
assumed to vary linearly over the triangular or tetrahedral elements of the mesh
(Jameson et al 1986, Rostand & Stoufflet 1988, Barth 1991a). Such a scheme
corresponds to a central difference discretization on structured meshes and
requires the addition of artificial diffusive terms to ensure stability. By analogy
with the blended second and fourth differences employed on structured meshes
(Jameson et al 1981), artificial dissipation is constructed as a combination of
an undivided Laplacian and biharmonic operator. First-order shock-capturing
diffusive terms are provided by the Laplacian operator, which is turned off away
from shocks in regions of smooth flow where the biharmonic dissipation ensures
stability with second-order accuracy. This approach has been used extensively
to compute inviscid and viscous two- and three-dimensional flows (Jameson
et al 1986; Mavriplis 1987, 1992; Smith 1990; Mavriplis & Venkatakrishnan
December 2, 1996 16:42 Annual Reviews Chapter14 AR23-14
490 MAVRIPLIS
by L and R in Figures 7 and 8), are then simply taken as the values in the neigh-
boring control volumes. For second-order approximations, the flow variables
are assumed to vary linearly over the control volumes, and interface values
must be extrapolated from centroidal values to the control-volume interfaces
using estimated gradient information. Higher-order methods may similarly be
L R
j
i
common being those of Osher (1984) and Roe (1981). In order to prevent os-
cillations near discontinuities such as shocks, the extrapolated interface values
must be limited, such that the creation of new extrema is never permitted.
Some of the earliest applications of upwind schemes to unstructured meshes
were reported by Desideri & Dervieux (1988). They developed a vertex-based
scheme that estimates gradients along an edge from the projection of gra-
by CAPES on 03/24/09. For personal use only.
492 MAVRIPLIS
i
dt
from the positivity of the Ci j . Thus maxima must be decreasing, and con-
versely minima must be increasing. This Local Extremum Diminishing (LED)
principle (Jameson 1994) represents a multidimensional generalization of the
Total Variation Diminishing (TVD) concept (Harten 1984). This has important
implications for the construction of robust and accurate schemes. The LED
by CAPES on 03/24/09. For personal use only.
principle can be used to guarantee the capture of shock waves without any os-
cillations. Furthermore, the maximum principle inherent in these schemes can
provide a robust solution strategy—for example, by guaranteeing the positivity
of all flow quantities simply if they are initialized as positive quantities.
Unfortunately, if the Ci j’s are constant, the scheme is linear and can be at
best first-order accurate. To achieve second-order accuracy, a nonlinear scheme
must be constructed, where the Ci j coefficients depend on the solution variables.
These usually include the use of limiters based on ratios of successive gradients
on neighboring mesh intervals. These methods have their roots in the early
work of Boris & Book (1973) on flux-corrected transport schemes. The original
idea is to construct a first-order scheme, the accuracy of which is then refined
by adding specific amounts of antidiffusive fluxes, under the constraints of
monotonicity. This scheme was extended to unstructured meshes by Löhner in
the FEM-FCT schemes (Löhner et al 1987). Jameson has derived a Symmetric
Limited Positive (SLIP) scheme that is both positive and second-order accurate
in smooth regions for unstructured meshes (Jameson et al 1986), where one-
dimensional solution variations along stencil edges are obtained by projecting
adjacent triangle gradients into the direction of the edge, in a manner similar to
that of Desideri & Dervieux (1988) (cf Figure 8). These schemes can be applied
to the unsplit form of the governing flow equations, in a scalar dissipation
mode, or to the split equations, characteristically or otherwise, to achieve more
accurate resolution of shock waves and boundary layers. However, for systems
of equations, monotonicity is not strictly guaranteed, as it is in the scalar case.
Additionally, these schemes rely on the use of a nonlinear limiter in all regions
of the flow-field (not only in the vicinity of shock waves), which may adversely
impact convergence, depending on the form of limiter employed.
The weak link in the upwind schemes described in the previous section in-
volves the use of a dimensionally split Riemann solver, which is applied in
December 2, 1996 16:42 Annual Reviews Chapter14 AR23-14
scalar advection equation have been developed and are compared in Paillere
et al (1993). Extensions of fluctuation-splitting schemes to systems of equa-
tions rely on the decomposition of the flux divergence into scalar waves. The
decomposition is not unique, and several strategies such as characteristic de-
composition (Deconinck et al 1986), simple wave models (Parpia & Michalak
1993, Roe 1986), and algebraic approaches (Sidilkover 1994) have been in-
vestigated. The advantages of fluctuation-splitting schemes are that they do
not require the use of dimensionally split Riemann solvers and they result in
a compact nearest-neighbor stencil for simplicial meshes. However, they are
second-order accurate only at steady state (Venkaktkrishnan 1996).
In the finite-element approach, the solution is expanded in a set of basis func-
tions, and the discrete equations are obtained by requiring the residuals to be
orthogonal to a set of trial functions. For second-order approximations, the basis
functions are generally taken as linear functions of compact support, which have
the value 1 at a given vertex and vanish at all other vertices. The solution expan-
sion is obtained by multiplying these basis functions by vertex-based-solution
variables and summing over the entire grid. When the test functions and the ba-
sis functions are the same, the method is called Galerkin, otherwise it is called
Petrov-Galerkin. The correspondence between (a) a Galerkin finite-element
strategy using linear elements and (b) finite-volume based central-differencing
has already been discussed. For the convective terms, additional artificial dis-
sipation is required for stability. The Galerkin finite-element method can also
be employed to discretize the viscous terms for the Navier-Stokes equations,
as discussed in the next section.
The Streamwise Upwind Petrov Galerkin (SUPG) (Hughes & Brooks 1979)
and Galerkin least-squares methods (Hughes et al 1989, Hughes 1987) intro-
duce dissipation to a regular Galerkin method by adding terms proportional
to the residual. In contrast to traditional artificial-dissipation formulations,
December 2, 1996 16:42 Annual Reviews Chapter14 AR23-14
494 MAVRIPLIS
the order of accuracy near discontinuities has been proposed by Hughes &
Mallet (1986).
496 MAVRIPLIS
each end of the edge. In order to compute this flux, a face area must be stored
for each edge, which corresponds to the area of the dual control-volume face
associated with the mesh edge in the vertex-based scheme, and to the area of
the cell face pierced by the dual edge in the cell-centered scheme. The use of
edge-based data structures results in lower memory overheads and increased
computational throughputs because redundant computations are eliminated and
the gather-scatter required for vectorization in supercomputers is minimized.
Annu. Rev. Fluid Mech. 1997.29:473-514. Downloaded from arjournals.annualreviews.org
Furthermore, because sets of edges can be used as building blocks for arbitrarily
shaped elements, hybrid meshes with mixed element types may be handled by
a single edge-based data structure, at least for inviscid flows.
For viscous flows, the Galerkin finite-element discretization of the diffusion
terms on simplicial meshes results in a nearest-neighbor stencil and thus may
be implemented using an edge-based data structure (Mavriplis 1995c). How-
by CAPES on 03/24/09. For personal use only.
ever, this is not the case for nonsimplicial meshes, since the resulting stencils
involve vertices that are not connected to the center vertex by a mesh edge
(such as diagonally opposed vertices in hexahedral elements). In these situ-
ations, the element-based data structure must be retained (Braaten & Connell
1996). An alternative is to resort to the thin-layer approximation of the vis-
cous terms on nonsimplicial meshes, which can be implemented exclusively
along edges (Mavriplis & Venkatakrishnan 1995b). The limitations of this
approach are obvious, although it is justifiable for highly stretched prismatic
or hexahedral meshes, where streamwise resolution has been sacrificed for
efficiency.
An interesting property of the edge-based data structure is that it can pro-
vide an interpretation of the discrete operator as a sparse matrix. For nearest-
neighbor stencil discretizations, all points in the stencil are joined to the center
point by a mesh edge. The discretization operator can be written as a sparse
matrix, where each nonzero entry in the matrix corresponds to a stencil coef-
ficient or edge of the mesh. For systems of equations, the edges correspond
to nonzero block matrix entries in the large sparse matrix. This interpretation
has implications for the implementation of implicit and algebraic multigrid
solution schemes (Venkatakrishnan & Mavriplis 1993, Mavriplis & Venkata-
krishnan 1995b).
One of the disadvantages of the edge-based data structure is that it requires
a preprocessing operation to extract a unique list of edges from the list of mesh
elements and to compute the associated edge coefficients. For unsteady flows
with dynamic meshes, this preprocessing must be performed every time the
mesh is altered, although this may be done locally. Additionally, for dynamic
grid cases, the element data structures are generally required for performing
mesh motion or adaptation, since edge lists represent a lower-level description
of the mesh.
December 2, 1996 16:42 Annual Reviews Chapter14 AR23-14
4. SOLUTION TECHNIQUES
The discretization of the spatial terms in the governing equations results in a
large coupled set of ordinary differential equations of the form
d(V Mw)
+ R(w) = 0 (2)
dt
where V represents the local control volume, M the mass matrix, and R(w)
Annu. Rev. Fluid Mech. 1997.29:473-514. Downloaded from arjournals.annualreviews.org
the spatially discretized terms. For a vertex-based scheme, the mass matrix
relates the average value of a control volume to the values at the vertices of
the mesh. Both V and M are constants for static meshes, and can therefore
be taken outside of the time derivative. For steady-state cases, these equations
must be integrated in time towards t → ∞, where the time derivatives become
vanishingly small. Since time accuracy is not a concern in such cases, the mass
by CAPES on 03/24/09. For personal use only.
498 MAVRIPLIS
∂w
represents the Jacobian and constitutes a large sparse matrix. At each
time step, the above linear system must be solved for the corrections 1w =
(wn+1 − wn ) from which the flow variables can be updated.
Implicit schemes may be classified by the degree to which the true Jacobian
matrix is approximated. If an exact linearization is employed, the method is
unconditionally stable and reduces to Newton’s method for 1t → ∞. Al-
by CAPES on 03/24/09. For personal use only.
500 MAVRIPLIS
the original Jacobian are retained, and the point-implicit method is obtained
(Hassan et al 1990, Thareja et al 1988). These terms represent the coupling be-
tween the various fluid dynamic equations at a grid point. While point-implicit
methods can offer increases in efficiency over regular explicit schemes, they
are also plagued by slow convergence for fine grids. Point-implicit meth-
ods are sometimes viewed as preconditioning techniques for explicit schemes,
by CAPES on 03/24/09. For personal use only.
where the point-implicit matrix is the preconditioner (Riemslagh & Dick 1993,
Morano & Dervieux 1993, Ollivier-Gooch 1995, Pierce & Giles 1996).
502 MAVRIPLIS
If the fine mesh is the product of an adaptive meshing algorithm using element
subdivision, then the nested parent meshes are natural candidates for coarse
multigrid levels (Parthasarathy & Kallinderis 1994, Braaten & Connell 1996).
Although this approach can be employed effectively when multiple adaptive
levels are present, the nested multigrid strategy places conflicting demands on
the multilevel construction. On the one hand, a very coarse mesh is desired
for rapid convergence, while a high-quality fine mesh is required for solution
Annu. Rev. Fluid Mech. 1997.29:473-514. Downloaded from arjournals.annualreviews.org
terpolation to transfer variables back and forth between the various meshes of
the sequence, within a multigrid cycle (Mavriplis 1992, 1995c; Peráire et al
1992; Leclercq 1990; Morano & Dervieux 1993; Riemslagh & Dick 1993).
The meshes may be generated using any feasible grid generation technique and
will generally be nonnested, and may even not contain any common points. The
only requirement is that they conform to the same domain boundaries. This
technique is more flexible than the nested subdivision approach, since the fine
and coarse meshes are not constrained, and may be optimized independently
for accuracy and speed of convergence respectively. Furthermore, this ap-
proach can be applied to a problem with a prespecified fine mesh. The intergrid
transfer operators can be determined in a preprocessing operation but require
smart search techniques to determine the patterns between coarse and fine grid
elements, which in principle may overlap randomly. Once these have been
determined and stored, grid transfers are simply implemented as a weighted
gather and scatter of data between coarse and fine grid arrays. Application of
this technique to a three-dimensional inviscid flow is depicted in Figures 10
and 11. The fine grid contains 804,000 points, or 4.5 million elements, and the
overset multigrid solver achieves a six-order reduction in the residual in approx-
imately 100 multigrid cycles, which represents close to an order of magnitude
savings over the single-grid scheme.
One of the drawbacks of this approach is that it still requires the user to
generate multiple grids for a single solution, which is at best tedious. With this
in mind, several efforts have been pursued to automate the process. Methods
that remove selected mesh points, using graph-based algorithms, and retriangu-
late the remaining subset of vertices have been demonstrated as effective tools
for generating coarse-level multigrid meshes (Guillard 1993, Chan & Smith
1993, Matheson & Tarjan 1994). However, for very complex geometries, the
December 2, 1996 16:42 Annual Reviews Chapter14 AR23-14
Figure 10 Two grids from the sequence of grids employed in the overset mesh multigrid scheme
for the computation of transonic flow over an aircraft configuration.
December 2, 1996 16:42 Annual Reviews Chapter14 AR23-14
504 MAVRIPLIS
Annu. Rev. Fluid Mech. 1997.29:473-514. Downloaded from arjournals.annualreviews.org
by CAPES on 03/24/09. For personal use only.
Figure 11 Computed solution displayed as Mach contours and observed convergence rates for
the single-grid, overset-multigrid, and agglomeration-multigrid algorithms for the computation of
transonic flow over an aircraft configuration.
December 2, 1996 16:42 Annual Reviews Chapter14 AR23-14
generation of coarse meshes that conform to the original geometry either man-
ually or automatically, becomes a difficult task.
An alternative is provided by agglomeration multigrid methods. Based on the
control-volume formulation, these methods form coarser-level meshes by fus-
ing or agglomerating fine-level control volumes together with their neighbors
(Lallemand et al 1992, Smith 1990, Venkatakrishnan & Mavriplis 1995a). The
resulting coarse mesh contains a smaller number of larger and more complex
Annu. Rev. Fluid Mech. 1997.29:473-514. Downloaded from arjournals.annualreviews.org
control volumes. The precise manner in which control volumes are agglom-
erated can be controlled through a graph-based algorithm similar to the vertex
removal procedure described above. The flow solver must be modified to run
on arbitrarily shaped control volumes on the coarse levels. For inviscid-flow
control-volume formulations, this presents little difficulty, since the equations
are generally discretized as fluxes over individual control-volume faces, which
by CAPES on 03/24/09. For personal use only.
can be used to build contour integrals about arbitrarily shaped control volumes.
Figure 12 depicts two coarse levels generated by the agglomeration algorithm
for the same case discussed above in the context of the overset-mesh multi-
grid approach. The observed convergence rate of the agglomeration multigrid
method is almost identical to that obtained with the overset-mesh method, as
depicted in Figure 11.
For viscous flows, the discretization of diffusion terms on arbitrarily shaped
control volumes is no longer straightforward. An algebraic interpretation of
agglomeration multigrid provides a mechanism for dealing with equations sets
that contain diffusion terms. Borrowing from the algebraic multigrid literature
(Ruge & Stüben 1987), a coarse-grid operator may be constructed by projecting
the fine-grid operator onto the space spanned by the coarse-grid basis functions.
If IhH represents the restriction or fine- to coarse-grid transfer operator, I Hh the
prolongation or coarse- to fine-grid operator, and A the fine-grid discretization
operator, then the so-called Galerkin coarse-grid operator is given by the se-
quence of operators IhH AI Hh . If A represents a finite-volume approximation
for inviscid flow on the fine grid, and restriction and prolongation are taken as
piecewise constant operators, then the Galerkin construction recovers the tradi-
tional coarse-grid finite-volume discretization discussed above. However, this
strategy can be extended in a straightforward manner for other types of govern-
ing equations, such as the Navier-Stokes equations and field-equation turbulent
models. This approach may be interpreted algebraically as summing the con-
stituent fine-grid equations within an agglomerated coarse-grid cell to form a
smaller number of coarse-grid equations. This algebraic interpretation allows
one to decouple the geometry from the solution process, thus discarding any
notion of coarse-level spatial discretizations, which enhances the robustness of
the solution process for complicated geometries. A solution of viscous turbu-
lent flow over a three-dimensional wing configuration using the agglomeration
December 2, 1996 16:42 Annual Reviews Chapter14 AR23-14
506 MAVRIPLIS
Annu. Rev. Fluid Mech. 1997.29:473-514. Downloaded from arjournals.annualreviews.org
by CAPES on 03/24/09. For personal use only.
Figure 12 Two coarse agglomerated levels from the sequence of levels employed by the
agglomeration-multigrid method for the computation of transonic flow over an aircraft config-
uration.
December 2, 1996 16:42 Annual Reviews Chapter14 AR23-14
and 14. The slower convergence in the viscous flow cases is principally due to
the anisotropic stretching of the mesh in the boundary-layer and wake regions.
Unstructured multigrid methods offer the possibility of designing schemes that
are insensitive to anisotropic effects. The graph-based agglomeration or vertex-
coarsening algorithms described above can be modified to merge only those
neighboring control volumes or delete only those neighboring vertices that are
by CAPES on 03/24/09. For personal use only.
508 MAVRIPLIS
Annu. Rev. Fluid Mech. 1997.29:473-514. Downloaded from arjournals.annualreviews.org
by CAPES on 03/24/09. For personal use only.
Figure 13 Initial 300,000-point grid and coarse agglomerated level employed in the agg-
lomeration-multigrid scheme for the computation of viscous turbulent flow over a wing config-
uration.
December 2, 1996 16:42 Annual Reviews Chapter14 AR23-14
Figure 14 Computed solution displayed as Mach contours on the wall and pressure contours
on the wing surface for fine 2.4 million–point (13.6 million–tetrahedra) grid and observed con-
vergence rates of agglomeration-multigrid algorithm on 300,000-point grid and 2.4 million–point
grid, demonstrating grid-independent convergence rates.
December 2, 1996 16:42 Annual Reviews Chapter14 AR23-14
510 MAVRIPLIS
While these strategies may yield desirable robustness characteristics, they are
plagued by the high memory requirements of the implicit linearization, which
are avoided in the FAS approach.
5. CONCLUSIONS
An overview of the state-of-the-art in the areas of mesh generation, adaptivity,
discretizations, and solution strategies for unstructured meshes has been given.
Annu. Rev. Fluid Mech. 1997.29:473-514. Downloaded from arjournals.annualreviews.org
An attempt has been made to outline the merits and disadvantages of the various
approaches in these areas. Other areas such as solution techniques for unsteady
flows and parallel implementations have not been addressed owing to space
limitations. Excellent survey papers in these areas, as well as more in-depth
reviews of some of the items covered in this paper, can be found in the literature
(Thompson & Weatherill 1993; Baker 1996; Venkatakrishnan 1995, 1996;
by CAPES on 03/24/09. For personal use only.
ACKNOWLEDGMENTS
Dr. V. Venkatakrishnan is acknowledged for his helpful comments in preparing
this paper. The author is also grateful to Dr. K. W. Anderson, Dr. S. Pirzadeh,
and E. Nielsen for providing figures for this paper.
Literature Cited
Aftosmis M, Gaitonde D, Tavares TS. 1994. structured meshes. AIAA Pap. 94-0415
On the accuracy, stability and monotonicity Anderson WK. 1994. A grid generation and
of various reconstruction algorithms for un- flow solution method for the Euler equa-
December 2, 1996 16:42 Annual Reviews Chapter14 AR23-14
tions on unstructured grids. J. Comput. Phys. Cai XC, Gropp WD, Keyes DE, Tidriri
110(1):23–38 MD. 1994. Newton-Krylov-Schwarz meth-
Anderson WK, Rausch R, Bonhaus D. 1995. ods in CFD. Proceedings of the International
Implicit multigrid algorithms for incom- Workshop on Numerical Methods for the
pressible turbulent flows on unstructured Navier-Stokes Equations. Notes on Numer-
grids. Proc. AIAA CFD Conf., 12th, San ical Fluid Mechanics, Vol. 47:17–30. Braun-
Diego. AIAA Pap. 95-1740-CP schweig/Wiesbaden: Vieweg
Babushka I, Aziz AK. 1976. On the angle con- Castro-Diaz MJ, Hecht F, Mohammadi B. 1995.
dition in the finite-element method. SIAM J. Anisotropic unstructured mesh adaptation for
Numer. Anal. 13(6) flow simulations. Int. Mesh Roundtable, 4th,
Baker TJ. 1987. Three dimensional mesh gener- Albuquerque, NM, pp. 73–85
Annu. Rev. Fluid Mech. 1997.29:473-514. Downloaded from arjournals.annualreviews.org
ation by triangulation of arbitrary point sets. Chan TF, Mathew TP. 1994. Domain decompo-
AIAA Pap. 87-1124 sition algorithms. Acta Numer. pp. 61–143
Baker TJ. 1991. Unstructured meshes and sur- Chan TF, Smith B. 1993. Domain decompo-
face fidelity for complex shapes. Proc. AIAA sition and multigrid algorithms for ellip-
CFD Conf., 10th, Honolulu, pp. 714–25. tic problems on unstructured meshes. UCLA
AIAA Pap. 91-1591-CP CAM Rep. 93-42, Dep. Math., Univ. Calif.,
Baker TJ. 1996. Mesh adaptation strategies for Los Angeles
problems in fluid dynamics. Finite Elem. Chew LP. 1989. Constrained Delaunay triangu-
Anal. Des. In press lations. Algorithmica 4:97–108
by CAPES on 03/24/09. For personal use only.
Baldwin BS, Barth TJ. 1991. A one-equation Connell SD, Braaten ME. 1995. Semi-
turbulence transport model for high Reynolds structured mesh generation for three-
number wall-bounded flows. AIAA Pap. 91- dimensional Navier-Stokes calculations.
0610 AIAA J. 33(6):1017–24
Barth TJ. 1991a. Numerical aspects of comput- Deconinck H, Hirsch Ch, Peuteman J. 1986.
ing viscous high Reynolds number flows on Characteristic decomposition methods for
unstructured meshes. AIAA Pap. 91-0721 the multidimensional Euler equations. Lect.
Barth TJ. 1991b. A three-dimensional upwind Notes Phys. 264:216–22
Euler solver for unstructured meshes. AIAA Desideri JA, Dervieux A. 1988. Compressible
Pap. 91-1548-CP flow solvers using unstructured grids. VKI
Barth TJ. 1992. Aspects of unstructured grids Lect. Ser. 1988-05:1–115
and finite-element volume solvers for the Eu- Edelsbrunner H, Tan TS, Waupotitsch R. 1992.
ler and Navier-Stokes equations. von Karman An 0(N 2 log N ) time algorithm for the min-
Inst. Lect. Ser., AGARD Publ. R-787 max angle triangulation. SIAM J. Sci. Stat.
Barth TJ, Jespersen DC. 1989. The design and Comput. 13:994–1008
application of upwind schemes on unstruc- Fezoui L, Stoufflet B. 1989. A class of im-
tured meshes. AIAA Pap. 89-0366 plicit upwind schemes for Euler simulations
Barth TJ, Linton SW. 1995. An unstructured with unstructured meshes. J. Comput. Phys.
mesh Newton solver for compressible fluid 84:174–206
flow and its parallel implementation. AIAA Frink NT. 1992. Upwind scheme for solving the
Pap. 95-0221 Euler equations on unstructured tetrahedral
Batina JT. 1993. Implicit upwind solution al- meshes. AIAA J. 30(1):70–77
gorithms for three-dimensional unstructured Frink NT. 1994. Recent progress toward a three-
meshes. AIAA J. 31(5):801–5 dimensional unstructured Navier-Stokes flow
Baum JD, Luo H, Löhner R. 1994. A new ALE solver. AIAA Pap. 94-0061
adaptive unstructured methodology for the George PL, Hecht F, Saltel E. 1991. Auto-
simulation of moving bodies. AIAA Pap. 94- matic mesh generator with specified bound-
0414 ary. Comput. Methods Appl. Mech. Eng.
Blacker TD, Stephenson MB. 1991. Paving: 33:975–95
a new approach to automated quadrilateral Guillard H. 1993. Node nested multigrid with
mesh generation. Int. J. Numer. Methods Eng. Delaunay coarsening. INRIA Rep. No. 1898
32:811–47 Gumbert C, Lohner R, Parikh P, Pirzadeh S.
Boris JP, Book DL. 1973. Flux corrected trans- 1989. A package for unstructured grid gen-
port, 1 SHASTA, a fluid transport algorithm eration and finite element flow solvers. AIAA
that works. J. Comput. Phys. 11:38–69 Pap. 89-2175
Bowyer A. 1981. Computing Dirichlet tessala- Harten A. 1984. On a class of high-reso-
tions. Comput. J. 24(2):162–66 lution total-variation-stable finite difference
Braaten ME, Connell SD. 1996. Three di- schemes. SIAM J. Numer. Anal. 21(1):1–
mensional unstructured adaptive multigrid 23
scheme for the Navier-Stokes equations. Hassan O, Morgan K, Peraire J. 1989. An adap-
AIAA J. 34(2):281–90 tive implicit/explicit finite element scheme
December 2, 1996 16:42 Annual Reviews Chapter14 AR23-14
512 MAVRIPLIS
for compressible high speed flows. AIAA Pap. JD Rice, pp. 161–95. New York: Academic
89-0363 Leclercq MP. 1990. Resolution des equations
Hassan O, Morgan K, Peraire J. 1990. An im- d’Euler par des methods multigrilles con-
plicit finite element method for high speed ditions aux limites en regime hypersonique.
flows. AIAA Pap. 90-0402 PhD thesis. Dep. Appl. Math, Univ. de Saint-
Holmes DG, Snyder DD. 1988. The generation Etienne
of unstructured meshes using Delaunay trian- Lo SH. 1985. A new mesh generation scheme
gulation. Numer. Grid Gener. Comput. Fluid for arbitrary planar domains. Int. J. Numer.
Mech. Proc. Int. Conf. Numer. Grid Gener. Methods Eng. 21:1403–26
Comput. Fluid Dyn., 2nd, Miami, ed. S Sen- Löhner R. 1993. Matching semi-structured and
gupta, J Hauser, PR Eisman, JF Thompson unstructured grids for Navier-Stokes calcula-
Annu. Rev. Fluid Mech. 1997.29:473-514. Downloaded from arjournals.annualreviews.org
Mavriplis DJ. 1992. Three-dimensional multi- laminar flows. Proc. 6th Int. Symp. CFD,
grid for the Euler equations. AIAA J. Lake Tahoe, NV
30(7):1753–61 Osher S. 1984. Riemann solvers, the en-
Mavriplis DJ. 1995a. Multigrid techniques for tropy condition and difference approxima-
unstructured meshes. VKI Lect. Ser. Comput. tions. SIAM J. Numer. Anal. 21(2):217–
Fluid Dyn., 26th, VKI-LS 1995-02 35
Mavriplis DJ. 1995b. A three-dimensional Paillere H, Deconinck H, Struijs R, Roe PL,
multigrid Reynolds-averaged Navier-Stokes Mesaros LM, Muller JD. 1993. Computa-
solver for unstructured meshes. AIAA J. tions of compressible flows using fluctuation-
33(3):445–53 splitting on triangular meshes. AIAA Pap. 93-
Mavriplis DJ. 1995c. Unstructured mesh gener- 3301-CP
Annu. Rev. Fluid Mech. 1997.29:473-514. Downloaded from arjournals.annualreviews.org
ation and adaptivity. VKI Lect. Ser. Comput. Palmerio B. 1994. An attraction-repulsion mesh
Fluid Dyn., 26th, VKI-LS 1995-02 adaption model for flow solution on un-
Mavriplis DJ, Venkatakrishnan V. 1994. Ag- structured grids. Comput. Fluids 23(3):487-
glomeration multigrid for two-dimensional 506
viscous flows. Comput. Fluids 24(5):553-70 Parpia I, Michalak DJ. 1993. Grid-independent
Mavriplis DJ, Venkatakrishnan V. 1995a. A upwind scheme for multidimensional flows.
3D agglomeration multigrid solver for the AIAA J. 31(4):646–51
Reynolds-averaged Navier-Stokes equations Parthasarathy V, Kallinderis Y. 1994. New
on unstructured meshes. AIAA Pap. 95-0345 multigrid approach for three-dimensional
by CAPES on 03/24/09. For personal use only.
514 MAVRIPLIS
mensional gas dynamics. J. Comput. Phys. upon a Voronoi type method. See Mavriplis
63:458–76 1991b, pp. 93–103
Rostand P, Stoufflet B. 1988. TVD schemes to van Leer B, Tai CH, Powell KG. 1989. Design
compute compressible viscous flows on un- of optimally-smoothing multi-stage schemes
structured meshes. Notes on Numerical Fluid for the Euler equations. AIAA Pap. 89-1933
Mechanics, Vol. 24:510–520. Braunschweig, Venkatakrishnan V. 1994. Parallel implicit
Ger: Vieweg unstructured grid Euler solvers. AIAA J.
Ruge JW, Stüben K. 1987. Algebraic multigrid. 32(10):1985–91
Multigrid Methods, SIAM Frontiers in Ap- Venkatakrishnan V. 1995a. Implicit schemes
plied Mathematics, ed. SF McCormick, pp. and parallel computing in unstructured grid
73–131. Philadelphia: SIAM CFD. VKI Lect. Ser. VKI-LS 1995-02
Annu. Rev. Fluid Mech. 1997.29:473-514. Downloaded from arjournals.annualreviews.org
Shakib F, Hughes TJR, Johan Z. 1989. A multi- Venkatakrishnan V. 1995b. On the accuracy of
element group preconditioned GMRES al- limiters and convergence to steady state so-
gorithm for nonsymmetric problems arising lutions. J. Comput. Phys. 118:120–30
in finite element analysis. Comput. Methods Venkatakrishnan V. 1996. A perspective on
Appl. Mech. Eng. 87:415–56 unstructured grid flow solvers. AIAA J.
Sidilkover D. 1994. A genuinely multidimen- 34(3):533–47
sional upwind scheme and efficient multigrid Venkatkrishnan V, Barth TJ. 1989. Application
solver for the compressible Euler equations. of direct solvers to unstructured meshes for
ICASE Rep. No. 94-84. ICASE, NASA Lan- the Euler and Navier-Stokes equations using
by CAPES on 03/24/09. For personal use only.
CONTENTS
G. I. TAYLOR IN HIS LATER YEARS, J. S. Turner 1